Wenlong Huangab,
Mingze Lvab,
Ying Li*ab,
Yushi Ding*ab,
Jiayao Luab,
Chunsheng Zhuangc,
Pengfei Yuec and
Wei Zhangc
aSchool of Metallurgy, Northeastern University, China. E-mail: liying@mail.neu.edu.cn; dingysh@smm.neu.edu.cn
bLiaoning Key Laboratory for Metallurgical Sensor Materials and Technology, Northeastern University, China
cInstitute of Applied Physics, Henan Academy of Sciences, Zhengzhou, Henan 450008, China
First published on 18th November 2024
Perovskite-type solid electrolytes exhibit a diverse range of conductive properties due to the competition and coupling of multiple degrees of freedom. In perovskite structures, B-site and X-site ions form topological octahedral sublattices, which are instrumental in regulating transport properties for various charge carriers. However, research focused on the relationship between octahedral distortion and conductive properties in perovskite-type proton conductors remains limited. In this study, dopants such as Ge, Sn, Pr, and Ce were selected to modify the degree of BO6 octahedral distortion in CaHf0.9Sc0.1O3−δ. The relationships between conductivity, transport number, mobility, and the distortion degree were systematically investigated. The data indicate that both proton and oxygen ion mobilities initially increase with the octahedral distortion angle and then decrease, and CaHf0.8Sn0.1Sc0.1O3−δ with an octahedral distortion angle of 15.6°, exhibited the highest ionic mobilities and conductivities. The BO6 octahedral distortion appears to limit oxide ion conduction while enhancing the proton transport number. However, excessive doping generates additional oxygen vacancies, which adversely affect proton conduction. Under the combined influence of these factors, CaHf0.8Ce0.1Sc0.1O3−δ achieved the highest proton transport number of 0.503 at 800 °C. Overall, this work provides insights into the relationship between octahedral distortion and conductive properties, suggesting that co-doping is a feasible approach for further regulating carrier mobility properties.
(1) |
Bonanos et al.9 reported that the orthorhombic distortion of perovskites can reduce oxide ion conductivity without significantly affecting protonic conductivity. Consequently, CaZrO3-type and CaHfO3-type proton conductors with orthorhombic distortion, particularly CaZr0.9Sc0.1O3−δ and CaHf0.9Sc0.1O3−δ, have shown the highest protonic transport numbers across various atmospheres.10–13 This phenomenon can be explained as follows:
In perovskite-type proton conductors, protonic migration primarily occurs via a rotation-jump mechanism. The energy barrier for this process is predominantly determined by the jump mechanism, with the jump barrier exhibiting a strong linear relationship with the distance between the proton and the second nearest oxygen (H–O).14 Similarly, the energy barrier for oxide ionic migration is influenced by the analogous O–O pathway.15 As a result, high protonic conductivity is often accompanied by high oxide ionic conductivity. Specifically, the tilting of the BO6 octahedra extends both the H–O and O–O distances.9,16 Given that oxide ions have a greater mass than protons, their transfer is slower under the same conditions. Therefore, the tilting of the BO6 octahedra hinders oxide ion movement more than proton movement, thereby increasing the protonic transport number at the cost of reduced overall conductivity.
Based on this mechanism, two methods for enhancing the protonic transport number can be readily inferred: (1) increasing the dopant element content M in AB1−xMxO3−δ, and (2) designing proton-conducting matrix materials with greater orthorhombic distortion, such as CaCeO3 and CaPrO3. However, at present, neither method achieves the desired results. In the first approach, increasing the concentration of trivalent dopant ions results in the formation of a significant number of oxygen vacancies in the perovskite structure, which leads to defect clustering and consequently reduces the conductivity of the proton conductor. In the second approach, the high degree of distortion in CaCeO3 and CaPrO3 exceeds the structural tolerance of perovskites, rendering their synthesis via conventional methods currently infeasible.
In our previous studies,17 we selected CaHf0.9Sc0.1O3−δ, which possesses a high protonic transport number, as the base material and utilized Ce, with its larger ionic radius compared to Hf, as a co-dopant to further enhance the proton transport number. This approach suggests that it is feasible to increase the orthorhombic distortion of the perovskite structure and thereby improve the proton transport number by co-doping with tetravalent elements without increasing oxygen vacancies. Nevertheless, we consider our prior research to be a preliminary exploration; thus, there remains a need to design a series of perovskites with varying degrees of distortion to establish a quantitative relationship between distortion degree and conductivity.
In this study, tetravalent dopants Sn, Ge, Pr, and Ce were selected to construct perovskites with varying degrees of distortion while minimizing changes to the oxygen vacancy concentration. Perovskites of the form CaHf0.9−xMxSc0.1O3−δ (M = Ge, Sn, Pr, and Ce; x = 0.1, 0.2) were prepared using the solid-state reaction sintering process. The valences of the dopant elements were determined by X-ray photoelectron spectroscopy (XPS). The conductivities of the resulting ceramics were measured at temperatures ranging from 400 °C to 800 °C using AC impedance spectroscopy. The partial conductivities and transport numbers of proton, oxide ion, and hole were analyzed through the defect equilibria model. The effects of dopant element and concentration on conductivities, ionic mobilities, and transport numbers were clarified. Additionally, a straightforward method for calculating the distortion angle of the BO6 octahedron based on the chemical formula was proposed. Common relationships between the distortion of the BO6 octahedron, the valence reduction tendencies, and the transport numbers and mobilities were clarified.
The crystal structures of the as-obtained CHMSs were characterized by X-ray diffraction (XRD, Rigaku Ultima IV), equipped with Cu Kα radiations (40 kV and 30 mA) in the 2θ range of 10–90°. The tolerance factors were calculated by eqn (2) using the Shannon ionic radii data.18
(2) |
The micro-morphologies of the ceramic specimens were observed by field emission scanning electron microscopy (FESEM, SU8010). The XPS spectra were recorded by ESCALAB 250Xi with a monochromatic Al/Mg binode X-ray source. The conductivities were determined by electrochemical impedance spectroscopy (EIS). To this end, both sides of sintered ceramic disks were first polished to ensure precise electrode area.
The surfaces were subsequently coated with Pt paste, connected with Pt electrode wires, and calcined at 900 °C for 30 min in air to create porous electrodes. A mixed atmosphere containing various partial pressures of oxygen and water vapor was flown through the samples at the rate of 100 mL min−1. The oxygen partial pressure was controlled using a mass flow controller, mixing O2 (99.9%), O2/Ar (5.00%), and Ar (99.9%). The water vapor partial pressure was controlled by evaporating quantitative water in a pre-furnace setup, while the water flow was controlled by a peristaltic pump with a 0.5 mm silicone tube.
The impedance spectroscopy of the samples was recorded by a frequency response analyzer (Solartron 1260, voltage 500 mV, measurement frequency 1 Hz to 5 MHz).
The partial conductivities of protons, oxide ions, and holes were analyzed by defect equilibria model,20–22 which can be expressed by eqn (3) and (4):
(3) |
(4) |
The relationship between the partial conductivities of protons, oxide ions, holes, and atmosphere can be calculated by equilibrium constant (eqn (3) and (4)):
(5) |
(6) |
(7) |
Transport numbers can be expressed as follows:
(8) |
The proton mobility and oxide ion mobility μO2− can be calculated as follows:
(9) |
(10) |
The proton concentration and oxide ion concentration cO2− can be calculated as follows:
(11) |
(12) |
Material | a (Å) | b (Å) | c (Å) | Unit cell volume (Å3) | Tolerance factor | Octahedra tilting (°) |
---|---|---|---|---|---|---|
CaZrO3 | 5.583 | 8.007 | 5.759 | 257.45 | 0.912 | 16.08 |
CaHf0.9Sc0.1O3−δ | 5.719 | 7.979 | 5.572 | 254.26 | 0.915 | 15.76 |
CaHf0.8Ge0.1Sc0.1O3−δ | 5.562 | 7.978 | 5.731 | 254.31 | 0.923 | 14.92 |
CaHf0.7Ge0.2Sc0.1O3−δ | 5.733 | 7.980 | 5.564 | 254.55 | 0.931 | 14.07 |
CaHf0.8Sn0.1Sc0.1O3−δ | 5.723 | 7.977 | 5.573 | 254.42 | 0.916 | 15.66 |
CaHf0.7Sn0.2Sc0.1O3−δ | 5.707 | 7.964 | 5.567 | 253.02 | 0.917 | 15.55 |
CaHf0.8Pr0.1Sc0.1O3−δ | 5.732 | 8.000 | 5.582 | 255.97 | 0.909 | 16.40 |
CaHf0.7Pr0.2Sc0.1O3−δ | 5.764 | 8.028 | 5.601 | 259.18 | 0.903 | 17.04 |
CaHf0.8Ce0.1Sc0.1O3−δ | 5.739 | 8.001 | 5.583 | 256.36 | 0.908 | 16.51 |
CaHf0.7Ce0.2Sc0.1O3−δ | 5.758 | 8.012 | 5.592 | 257.98 | 0.901 | 17.25 |
The tolerance factors, which indicate the distortion of the ABX3 lattice, can be calculated using eqn (2). However, the tolerance factor alone does not adequately represent the distortion associated with the tilting angle of the BO6 octahedra. For example, O'Keeffe24 calculated this tilting angle using eqn (13).
(13) |
However, numerous errors may be present in the calculation of fractional coordinates via XRD and Rietveld refinement. Additionally, doped perovskites often contain more than two types of BO6 octahedra, making the calculation of the tilting angle using this method challenging. Consequently, our study utilized CIF files of undoped ABX3-type oxides and halides from the Crystallography Open Database (COD). The angles between the BX6 octahedra and the lattice were measured using a geometric method in the crystal visualization software VESTA25 (Fig. 2). The relationship between the BO6 octahedral tilting and tolerance factor, depicted in Fig. 3, showed a linear increase in BX6 octahedral tilting as the tolerance factor decreased from 0.95 to 0.80. Notably, the crystal structure transitions to a cubic form when the tolerance factor exceeds 0.95. From this analysis, an equation was derived through curve fitting (eqn (14)):
φ = −106.05t + 112.80 | (14) |
Accordingly, the proposed perovskites, such as CHPS20 possessed HfO6, PrO6, and ScO6 types BO6 octahedra, with tilting angles of 15.61°, 21.61°, and 17.19°, respectively. The average tilting angle of CHPS20 estimated by eqn (10) was 17.04°.
Fig. 5 FESEM images of surface CHMSs: (a and b) M Ge, (c and d) M Sn, (e and f) M Pr, and (g and h) M Ce. (a, c, e, g) x = 0.1, (b, d, f, h) x = 0.2. |
Fig. 6 Nyquist plots of CHSS10: (a and b) under an atmosphere of 99.9% O2 and 2.34 kPa H2O at 400–800 °C, (c) at 500 °C, and (d) low-frequency parts versus different atmospheres at 600 °C. |
Fig. 6(d) shows the low-frequency region of Nyquist plots versus different atmospheres at 600 °C. As oxygen and vapor partial pressure rose, total resistance decreased with the lowest resistance observed under 2.34 kPa H2O and 99.9% O2. According to eqn (3) and (4), water and oxygen molecules competed to occupy the oxygen vacancies under wet atmosphere containing oxygen.37 Oxygen vacancies provided sites for the transport position of oxide ions and hydration, while oxidation decrease the concentration of oxygen vacancies and limit the oxide ionic conduction.
As a result, the oxide ionic transport number was always below 0.1 under similar wet atmosphere containing oxygen. Additionally, water increases the concentration of protons, while oxygen enhances the concentration of holes. Here, total resistances decreased by 11% with a 1.71% increase in water vapor, and resistance decreased by 6.2% with a 97% increase in water vapor. Thus, water molecules preferentially occupy oxygen vacancies, generating a higher concentration of protons than holes. Since mobility of proton was also higher than that of hole in ABO3-type proton conductor,37,38 protonic conduction likely dominates at 600 °C.
In general, such behavior is commonly observed in ABO3-type proton conductors, where the slope tends to increase in double-type proton conductors. This phenomenon may be attributed to the rise in protonic conductivity combined with a decline in oxide ionic conductivity as a function of vapor partial pressure (eqn (3)). The difference change in these conductivities determines the slope of conductivity versus . Therefore, the slope decreased with high protonic transport number, while it increases in proton conductors with significant oxide ionic conduction. Also, the oxide ionic transport numbers in AB′B′′O3 are often higher than those of simple ABO3 due to enhanced oxygen vacancy concentration tolerance in AB′B′′O3.38,39 In this study, all slopes decreased as a function of vapor partial pressure, likely due to high protonic transport number in CHMSs.
After determining the key parameter α of defect equilibria model in Fig. 7, the dependence of equilibrium constant (KH2O) values of eqn (3) on temperature of CHSSs were calculated (Fig. 8). All standard molar hydration enthalpy ΔHθm values of CHMSs were then calculated from the slope of ln KH2O versus temperature (Table 2).
BaZr0.8Y0.2O3−δ (ref. 40) | BaCe0.9Y0.1O3−δ (ref. 41) | CHGS20 | CHGS10 | CHSS20 | CHSS10 | CHPS10 | CHCS10 | CHPS20 | CHCS20 | |
---|---|---|---|---|---|---|---|---|---|---|
σtot (800 °C)/S cm−1 | ∼5 × 10−2 (700 °C) | ∼5 × 10−2 | 1.20 × 10−5 | 1.63 × 10−5 | 8.36 × 10−4 | 1.04 × 10−3 | 6.79 × 10−5 | 8.41 × 10−5 | 6.52 × 10−7 | 6.18 × 10−5 |
(800 °C)/S cm−1 | ∼8 × 10−3 | 5.19 × 10−6 | 7.64 × 10−6 | 3.67 × 10−4 | 4.99 × 10−4 | 3.38 × 10−5 | 4.23 × 10−5 | 2.83 × 10−7 | 2.67 × 10−5 | |
σO2− (800 °C)/S cm−1 | ∼3 × 10−2 | 3.06 × 10−6 | 3.39 × 10−6 | 2.08 × 10−4 | 1.99 × 10−4 | 9.69 × 10−6 | 1.16 × 10−5 | 1.18 × 10−7 | 1.19 × 10−5 | |
σh˙ (800 °C)/S cm−1 | ∼1 × 10−2 | 3.06 × 10−6 | 5.23 × 10−6 | 2.61 × 10−4 | 3.46 × 10−4 | 2.45 × 10−5 | 3.01 × 10−5 | 2.51 × 10−7 | 2.31 × 10−5 | |
ΔHθm/kJ mol−1 | −62.4 | −61.0 | −64.8 | −61.0 | −63.2 | −58.9 | −82.6 | −77.5 | ||
Etot/eV | ∼0.62 | 1.04 | 1.04 | 0.88 | 0.81 | 0.97 | 0.70 | 1.00 | 0.69 | |
EOH˙/eV | 0.94 | 0.94 | 0.78 | 0.72 | 0.89 | 0.62 | 0.90 | 0.59 | ||
EO2−/eV | 1.62 | 1.89 | 1.77 | 1.57 | 1.45 | 1.26 | 1.68 | 1.12 | ||
Eh˙/eV | 1.35 | 1.38 | 1.21 | 1.18 | 1.23 | 1.02 | 1.31 | 1.05 | ||
tOH˙ (800 °C) | <0.1 | ∼0.16 | 0.434 | 0.470 | 0.439 | 0.478 | 0.497 | 0.503 | 0.433 | 0.433 |
tO2− (800 °C) | ∼0.64 | 0.256 | 0.208 | 0.249 | 0.190 | 0.143 | 0.138 | 0.181 | 0.193 | |
th˙ (800 °C) | ∼0.20 | 0.310 | 0.322 | 0.312 | 0.331 | 0.360 | 0.358 | 0.385 | 0.373 | |
μOH˙/(cm2 (V−1 s−1))(800 °C) | ∼9.3 × 10−5 | ∼3.1 × 10−4 | 3.12 × 10−7 | 6.15 × 10−7 | 2.03 × 10−5 | 3.84 × 10−5 | 2.95 × 10−6 | 2.13 × 10−6 | 1.59 × 10−8 | 1.58 × 10−6 |
μO2−/(cm2 (V−1 s−1))(800 °C) | ∼1.3 × 10−7 | ∼1.4 × 10−3 | 2.41 × 10−7 | 2.93 × 10−7 | 8.96 × 10−6 | 1.13 × 10−5 | 4.96 × 10−7 | 4.86 × 10−7 | 8.63 × 10−9 | 7.71 × 10−7 |
cOH˙/mol% (800 °C) | ∼3.0 | 6.1 | 4.0 | 4.3 | 4.2 | 3.7 | 5.9 | 5.7 | 5.3 | |
cO2−/mol% (800 °C) | ∼3.5 | 6.9 | 12.1 | 10.2 | 11.6 | 12.7 | 8.1 | 8.7 | 9.4 |
In this study, the electrical properties are the primary focus of analysis, particularly the impact of the degree of octahedral distortion on these properties. Therefore, in Table 2, the materials are arranged differently from Table 1, being listed in the increasing order of distortion of BO6. The study was conducted under a common atmospheric condition (20% O2 and 2.34 kPa H2O/Ar), ensuring that all conduction properties were evaluated under the same environment. All CHMSs displayed ΔHθm values ranging from −58.9–−82.6 kJ mol−1. Løken et al.42 estimated ΔHθm values of CaZr0.9Sc0.1O3−δ and CaZr0.8Sc0.2O3−δ by TG-DSC method, and recorded −58 kJ mol−1 and −94 kJ mol−1, respectively. These values are comparable to those obtained in the current study. Note that ΔHθm represents the water absorbing power of proton conductors, which can be related to oxygen vacancy through eqn (3). The valence reduction of B-site ions can generate additional oxygen vacancies, with M4+ ions exhibiting higher valence reduction tendencies than Hf4+. Therefore, M4+ doped materials possessed more oxygen vacancy concentrations of CHS10, and all CHMS20 showed more negative ΔHθm values than CHMS10.
In addition, Pr4+ and Ce4+ have markedly higher valence reduction tendencies than Ge4+ and Sn4+, resulting in CHPS20 and CHCS20 having the most negative ΔHθm in Table 2. Hydration decreased the concentration of oxygen vacancy and limited the oxide ionic conduction, thereby increasing the protonic transport numbers and decreasing the oxide ionic transport numbers for CHPS20 and CHCS20 relative to other materials, considering the effects of hydration. However, more oxygen vacancy concentrations may negatively affect protonic conduction. Regarding the protonic microscopic migration pathway, proton can jump to nearest eight oxygen sites in an undoped or fully hydrated proton conductor. The number of jumpable oxygen sites decreases in proton conductors with more oxygen vacancies in the same atmosphere, thereby enhancing apparent activation energy. Therefore, the properties of CHMSs containing valence reduction elements warrant more comprehensive studies.
The partial conductivities of CHMSs under pH2O = 2.34 kPa and 20% O2/Ar were calculated by defect equilibria model, with select data presented in Table 2. Meanwhile, the partial conductivities and activation energies for each carrier of CHSSs are provided in Fig. 9. According to Table 2, CHSS displayed the highest conductivities, while an excess Pr dopant significantly restricted conduction due to the surplus of oxygen vacancies. Lim et al.21,22,43 estimated typical activation energies of proton, oxide ion, and hole to be ∼0.7 eV, ∼0.9 eV, and ∼1.4 eV, respectively. BaCe0.9Y0.1O3−δ (BCY10), like CHMSs, also possesses an orthorhombic structure with a pnma space group, but exhibits considerably less octahedral distortion relative to CHMSs. In contrast, the cubic structure of BaZr0.8Y0.2O3−δ (BZY20) lacks orthorhombic distortion. Consequently, the total activation energy of BCY10 is 0.62 eV, markedly lower than typical activation energy values. Furthermore, both the conductivities and protonic mobilities of BZY20 and BCY10 surpass those of CHMSs, implying that charge carrier conduction in these materials is not impeded by their structural configurations. In our study, the activation energies for the three carriers were calculated as ∼0.8 eV, ∼1.5 eV, and ∼1.2 eV, respectively. All activation energies of oxide ions were higher than those of holes, deviating from typical values. Hence, the conduction of oxide ions was strongly limited in CHMSs. Additionally, both the conduction of proton and oxide ions were limited, as expected from the distortion of BO6 octahedron. The activation energies of holes were lower than typical values owing to the additional oxygen vacancies generated by valence reduction. All protonic activation energies were significantly lower than those of oxide ions and holes, indicating CHMSs dominated by protonic conduction.
Transport numbers were calculated from Fig. 9 by eqn (8), and the values of CHSSs are depicted in Fig. 10. Transport properties in Table 2 indicate that all CHMSs dominated by protonic conduction at 400–800 °C. The proton transport numbers for BZY20 and BCY10 were notably lower than those for CHMSs, while their oxide ion transport numbers were the highest. This suggests that the orthorhombic distortion in CHMSs effectively restricts oxide ion conduction. Besides, transport numbers of CHMS10 were higher than those of CHMS20, indicating limited benefit of protonic conduction by excess oxygen vacancies generated by valence reduction. CHCS10 displayed the highest protonic transport number reaching 0.961–0.503 at 400–800 °C. Protonic transport numbers of CHMS10 increased in the following order: CHGS10 < CHSS10 < CHPS10 < CHCS10. Also, oxide ionic transport numbers of CHMS10 decreased in the same order. Both trends were closely aligned with the distortion degree of BO6 octahedron. Transport numbers of holes for CHMSs appeared similar, with slight increasing tendency following valence reduction. Thus, the distortion degree of BO6 octahedron effectively enhances protonic transport number.
It is worth noting that the transport number is a composite value, integrating various conductivities, and may not precisely elucidate the impact of distortion on distinct charge carriers. To further analyze the impact of octahedral distortion on proton and oxygen ion migration, the mobility and carrier concentration of both proton and oxygen ion were calculated (Table 2). As shown in Fig. 11(a), the logarithm of mobility exhibits a linear relationship with temperature. As temperature increases, ions acquire greater kinetic energy, resulting in increased mobility for both proton and oxygen ion. However, elevated temperatures adversely affect hydration (Fig. 8), leading to a gradual decrease in proton concentration with increasing temperature. Additionally, according to eqn (3), weakened hydration also results in a higher concentration of oxygen vacancies in the proton conductor (Fig. 11(b)). Fig. 11(c) illustrates the effect of octahedral distortion angle on proton conductor mobility. Notably, CHPS20 exhibits significantly different mobility compared to other materials, with both proton and oxygen ion being significantly suppression. We hypothesize that the excessive oxygen vacancies in CHPS20 may decrease the number of jumpable oxygen paths, thereby reducing the ionic mobility, which explains why CHPS20 does not align with the behavior of other materials. Interestingly, Fig. 11(c) shows that both proton and oxygen ion mobility initially increase with octahedral distortion angle and then decrease. The optimal octahedral distortion angle for ionic mobility is 15.6°, corresponding to the highest conductivity observed in CHSS10. Fig. 11(d) demonstrates that proton and oxygen ion concentrations remain unaffected by octahedral distortion; at 800 °C, the oxygen ion concentration in all materials exceeds the proton concentration. Overall, octahedral distortion affects ion mobility, ultimately influencing the conductivity of proton conductor.
This journal is © The Royal Society of Chemistry 2024 |